content
stringlengths 15
7.12M
| publicationDate
stringlengths 10
10
| abstract
stringlengths 41
2.5k
| title
stringlengths 14
230
| id
stringlengths 11
12
|
---|---|---|---|---|
Another view on Gilbert damping in two-dimensional ferromagnets Anastasiia A. Pervishko1, Mikhail I. Baglai1,2, Olle Eriksson2,3, and Dmitry Yudin1 1ITMO University, Saint Petersburg 197101, Russia 2Department of Physics and Astronomy, Uppsala University, Box 516, SE-75 121 Uppsala, Sweden 3School of Science and Technology, ¨Orebro University, SE-701 82 ¨Orebro, Sweden ABSTRACT A keen interest towards technological implications of spin-orbit driven magnetization dynamics requests a proper theoretical description, especially in the context of a microscopic framework, to be developed. Indeed, magnetization dynamics is so far approached within Landau-Lifshitz-Gilbert equation which characterizes torques on magnetization on purely phenomenological grounds. Particularly, spin-orbit coupling does not respect spin conservation, leading thus to angular momentum transfer to lattice and damping as a result. This mechanism is accounted by the Gilbert damping torque which describes relaxation of the magnetization to equilibrium. In this study we work out a microscopic Kubo-St ˇreda formula for the components of the Gilbert damping tensor and apply the elaborated formalism to a two-dimensional Rashba ferromagnet in the weak disorder limit. We show that an exact analytical expression corresponding to the Gilbert damping parameter manifests linear dependence on the scattering rate and retains the constant value up to room temperature when no vibrational degrees of freedom are present in the system. We argue that the methodology developed in this paper can be safely applied to bilayers made of non- and ferromagnetic metals, e.g., CoPt. Introduction In spite of being a mature field of research, studying magnetism and spin-dependent phenomena in solids still remains one of the most exciting area in modern condensed matter physics. In fact, enormous progress in technological development over the last few decades is mainly held by the achievements in spintronics and related fields1–11. However the theoretical description of magnetization dynamics is at best accomplished on the level of Landau-Lifshitz-Gilbert (LLG) equation that characterizes torques on the magnetization. In essence, this equation describes the precession of the magnetization, mmm(rrr;t), about the effective magnetic field, HHHeff(rrr;t), created by the localized moments in magnetic materials, and its relaxation to equilibrium. The latter, known as the Gilbert damping torque12, was originally captured in the form ammm¶tmmm, where the parameter adetermines the relaxation strength, and it was recently shown to originate from a systematic non-relativistic expansion of the Dirac equation13. Thus, a proper microscopic determination of the damping parameter a(or, the damping tensor in a broad sense) is pivotal to correctly simulate dynamics of magnetic structures for the use in magnetic storage devices14. From an experimental viewpoint, the Gilbert damping parameter can be extracted from ferromagnetic resonance linewidth measurements15–17or established via time-resolved magneto-optical Kerr effect18, 19. In addition, it was clearly demonstrated that in bilayer systems made of a nonmagnetic metal (NM) and a ferromagnet material (FM) the Gilbert damping is drastically enhanced as compared to bulk FMs20–24. A strong magnetocrystalline anisotropy, present in CoNi, CoPd, or CoPt, hints unambiguously for spin-orbit origin of the intrinsic damping. A first theoretical attempt to explain the Gilbert damping enhancement was made in terms of sdexchange model in Ref.25. Within this simple model, magnetic moments associated with FM layer transfer angular momentum via interface and finally dissipate. Linear response theory has been further developed within free electrons model26, 27, while the approach based on scattering matrix analysis has been presented in Refs.28, 29. In the latter scenario spin pumping from FM to NM results in either backscattering of magnetic moments to the FM layer or their further relaxation in the NM. Furthermore, the alternative method to the evaluation of the damping torque, especially in regard of first-principles calculations, employs torque-correlation technique within the breathing Fermi surface model30. While a direct estimation of spin-relaxation torque from microscopic theory31, or from spin-wave spectrum, obtained on the basis of transverse magnetic field susceptibility32, 33, are also possible. It is worth mentioning that the results of first-principles calculations within torque-correlation model34–38and linear response formalism39, 40reveal good agreement with experimental data for itinerant FMs such as Fe, Co, or Ni and binary alloys. Last but not least, an intensified interest towards microscopic foundations of the Gilbert parameter ais mainly attributed to the role the damping torque is known to play in magnetization reversal41. In particular, according to the breathing Fermi surface model the damping stems from variations of single-particle energies and consequently a change of the Fermi surfacearXiv:1807.07897v2 [cond-mat.mes-hall] 21 Nov 2018z yx FM NMFigure 1. Schematic representation of the model system: the electrons at the interface of a bilayer, composed of a ferromagnetic (FM) and a nonmagnetic metal (NM) material, are well described by the Hamiltonian (1). We assume the magnetization of FM layer depicted by the red arrow is aligned along the zaxis. shape depending on spin orientation. Without granting any deep insight into the microscopic picture, this model suggests that the damping rate depends linearly on the electron-hole pairs lifetime which are created near the Fermi surface by magnetization precession. In this paper we propose an alternative derivation of the Gilbert damping tensor within a mean-field approach according to which we consider itinerant subsystem in the presence of nonequilibrium classical field mmm(rrr;t). Subject to the function mmm(rrr;t)is sufficiently smooth and slow on the scales determined by conduction electrons mean free path and scattering rate, the induced nonlocal spin polarization can be approached within a linear response, thus providing the damping parameter due to the itinerant subsystem. In the following, we provide the derivation of a Kubo-St ˇreda formula for the components of the Gilbert damping tensor and illustrate our approach for a two-dimensional Rashba ferromagnet, that can be modeled by the interface between NM and FM layers. We argue that our theory can be further applied to identify properly the tensorial structure of the Gilbert damping for more complicated model systems and real materials. Microscopic framework Consider a heterostructure made of NM with strong spin-orbit interaction covered by FM layer as shown in Fig. 1, e.g., CoPt. In general FMs belong to the class of strongly correlated systems with partially filled dorforbitals which are responsible for the formation of localized magnetic moments. The latter can be described in terms of a vector field mmm(rrr;t)referred to as magnetization, that in comparison to electronic time and length scales slowly varies and interacts with an itinerant subsystem. At the interface (see Fig. 1) the conduction electrons of NM interact with the localized magnetic moments of FM via a certain type of exchange coupling, sdexchange interaction, so that the Hamiltonian can be written as h=p2 2m+a(sssppp)z+sssMMM(rrr;t)+U(rrr); (1) where first two terms correspond to the Hamiltonian of conduction electrons, on condition that the two-dimensional momentum ppp= (px;py) =p(cosj;sinj)specifies electronic states, mis the free electron mass, astands for spin-orbit coupling strength, while sss= (sx;sy;sz)is the vector of Pauli matrices. The third term in (1) is responsible for sdexchange interaction with the exchange field MMM(rrr;t) =Dmmm(rrr;t)aligned in the direction of magnetization and Ddenoting sdexchange coupling strength. We have also included the Gaussian disorder, the last term in Eq. (1), which represents a series of point-like defects, or scatterers, hU(rrr)U(rrr0)i= (mt) 1d(rrr rrr0)with the scattering rate t(we set ¯h=1throughout the calculations and recover it for the final results). Subject to the norm of the vector jmmm(rrr;t)j=1remains fixed, the magnetization, in broad terms, evolves according to (see, e.g., Ref.42), ¶tmmm=fffmmm=gHHHeffmmm+csssmmm; (2) where fffcorresponds to so-called spin torques. The first term in fffdescribes precession around the effective magnetic field HHHeffcreated by the localized moments of FM, whereas the second term in (2) is determined by nonequilibrium spin density of conduction electrons of NM at the interface, sss(rrr;t). It is worth mentioning that in Eq. (2) the parameter gis the gyromagnetic ratio, while c= (gmB=¯h)2m0=dis related to the electron g factor ( g=2), the thickness of a nonmagnetic layer d, with mBand m0standing for Bohr magneton and vacuum permeability respectively. Knowing the lesser Green’s function, G<(rrrt;rrrt), one can easily evaluate nonequilibrium spin density of conduction electrons induced by slow variation of magnetization orientation, sm(rrr;t) = i 2Tr smG<(rrrt;rrrt) =Qmn¶tmn+:::; (3) 2/8where summation over repeated indexes is assumed ( m;n=x;y;z). The lesser Green’s function of conduction electrons can be represented as G<= GK GR+GA =2, where GK,GR,GAare Keldysh, retarded, and advanced Green’s functions respectively. Kubo-St ˇreda formula We further proceed with evaluating Qmnin Eq. (3) that describes the contribution to the Gilbert damping due to conduction electrons. In the Hamiltonian (1) we assume slow dynamics of the magnetization, such that approximation MMM(rrr;t) MMM+(t t0)¶tMMMwith MMM=MMM(rrr;t0)is supposed to be hold with high accuracy, H=p2 2m+a(sssppp)z+sssMMM+U(rrr)+(t t0)sss¶tMMM; (4) where first four terms in the right hand side of Eq. (4) can be grouped into the Hamiltonian of a bare system, H0, which coincides with that of Eq. (1), provided by the static magnetization configuration MMM. In addition, the expression (4) includes the time-dependent term V(t)explicitly, as the last term. In the following analysis we deal with this in a perturbative manner. In particular, the first order correction to the Green’s function of a bare system induced by V(t)is, dG(t1;t2) =Z CKdtZd2p (2p)2gppp(t1;t)V(t)gppp(t;t2); (5) where the integral in time domain is taken along a Keldysh contour, while gppp(t1;t2) =gppp(t1 t2)[the latter accounts for the fact that in equilibrium correlation functions are determined by the relative time t1 t2] stands for the Green’s function of the bare system with the Hamiltonian H0in momentum representation. In particular, for the lesser Green’s function at coinciding time arguments t1=t2t0, which is needed to evaluate (3), one can write down, dG<(t0;t0) =i 2¥Z ¥de 2pZd2p (2p)2n gR pppsm¶g< ppp ¶e ¶gR ppp ¶esmg< ppp+g< pppsm¶gA ppp ¶e ¶g< ppp ¶esmgA pppo ¶tMm; (6) where m=x;y;z, while gR,gA, and g<are the bare retarded, advanced, and lesser Green’s functions respectively. To derive the expression (6) we made use of Fourier transformation gppp=Rd(t1 t2)gppp(t1 t2)eie(t1 t2)and integration by parts. To finally close up the derivation we employ the fluctuation-dissipation theorem according to which g<(e) = [gA(e) gR(e)]f(e), where f(e) = [eb(e m)+1] 1stands for the Fermi-Dirac distribution with the Fermi energy m. Thus, nonequi- librium spin density of conduction electrons (3) within linear response theory is determined by Qmn=Q(1) mn+Q(2) mn, where Q(1) mn=1 4Trh sm¥Z ¥de 2pZd2p (2p)2n¶gR ppp ¶esngR ppp gR pppsn¶gR ppp ¶e+gA pppsn¶gA ppp ¶e ¶gA ppp ¶esngA pppo f(e)i ; (7) which involves the integration over the whole Fermi sea, and Q(2) mn=1 4Trh sm¥Z ¥de 2p ¶f(e) ¶eZd2p (2p)2n gR pppsngR ppp+gA pppsngA ppp 2gR pppsngA pppoi ; (8) which selects the integration in the vicinity of the Fermi level. Generally, the form of Qmnbelongs to the class of Kubo-St ˇreda formula, and, in essence, represents the response to the external stimulus in the form of ¶tMn. We can immediately establish a quantitative agreement between the result given by Eq. (8) and the previous studies within a Kubo formalism40, 43–46which allow a direct estimation within the framework of disordered alloys. Formally, the expression (7) corresponds to the so-called St ˇreda contribution. Such a term was originally identified in Ref.47when studying quantum-mechanical conductivity. Notably, in Eq. (7) each term represents the product of either retarded or advanced Green’s functions. In this case the poles of the integrand function are positioned on the same side of imaginary plane, making disorder correction smaller in the weak disorder limit (see, e.g., Ref.48). Meanwhile, having no classical analog this contribution appears to be important enough when the spectrum of the system is gapped and the Fermi energy is placed exactly in the gap47. It is worth mentioning that the contribution due to Eq. (7) has never been discussed in this context before. In the meantime, Kubo-St ˇreda expression for the components of the Gilbert damping tensor has been addressed from the perspective of first-principles calculations49and current-induced torques50. 3/8Results and discussion Let us apply the formalism developed in the previous section to a prototypical model: we work out the Gilbert damping tensor for a Rashba ferromagnet with the magnetization mmm=zzzaligned along the zaxis. In the limit of weak disorder the Green’s function of a bare system can be expressed as gR ppp(e) = e H0 SR 1=e eppp+id+a(sssppp)z+(D+ih)sz (e eppp+id)2 a2p2 (D+ih)2; (9) where eppp=p2=(2m)is the electron kinetic energy. We put the self-energy SRdue to scattering off scalar impurities into Eq. (9), which is determined from SR= i(d hsz)(see, e.g., Ref.51). In particular, for jej>jDjwe can establish that d=1=(2t) andh=0 in the weak disorder regime to the leading order. Without loss of generality, in the following we restrict the discussion to the regime m>jDj, which is typically satisfied with high accuracy in experiments. As previously discussed, the contribution owing to the Fermi sea, Eq. (7), can in some cases be ignored, while doing the momentum integral in Eq. (8) results in, 1 mtZd2p (2p)2gR ppp(e)sssgA ppp(e) =D2 D2+2ersss+Dd D2+2er(ssszzz)+D2 er D2+2er(ssszzz)zzz; (10) where r=ma2. Thus, thanks to the factor of delta function d(e m) = ¶f(e)=¶e, to estimate Q(2) mnat zero temperature one should put e=min Eq. (10). As a result, we obtain, Q(2) mn= 1 4pm D2+2mr0 @2tmr D 0 D 2tmr 0 0 0 2 tD21 A: (11) Meanwhile, to properly account the correlation functions which appear when averaging over disorder configuration one has to evaluate the so-called vertex corrections, which from a physical viewpoint makes a distinction between disorder averaged product of two Green’s function, hgRsngAidis, and the product of two disorder averaged Green’s functions, hgRidissnhgAidis, in Eq. (8). Thus, we further proceed with identifying the vertex part by collecting the terms linear in dexclusively, GGGs=Asss+B(ssszzz)+C(ssszzz)zzz; (12) provided A=1+D2=(2er),B= (D2+2er)Dd=(D2+er)2, and C=D2=(2er) er=(D2+er). To complete our derivation we should replace snin Eq. (8) by Gs nand with the aid of Eq. (10) we finally derive at e=m, Q(2) mn=0 @Qxx Qxy 0 QxyQxx 0 0 mtD2=(4pmr)1 A: (13) We defined Qxx= mtmr=[2p(D2+mr)]andQxy= mD(D2+2mr)=[4p(D2+mr)2], which unambiguously reveals that account of vertex correction substantially modifies the results of the calculations. With the help of Eqs. (3), (11), and (13) we can write down LLG equation. Slight deviation from collinear configurations are determined by xandycomponents ( mxand myrespectively, so that jmxj;jmyj1). The expressions (11) and (13) immediately suggest that the Gilbert damping at the interface is a scalar, aG, ¶tmmm=˜gHHHeffmmm+aGmmm¶tmmm; (14) where the renormalized gyromagnetic ratio and the damping parameter are, ˜g=g 1+cDQxy;aG= cDQxx 1+cDQxy cDQxx: (15) In the latter case we make use of the fact that mc1for the NM thickness d100mm — 100 nm. In Eq. (14) we have redefined the gyromagnetic ratio g, but we might have renormalized the magnetization instead. From physical perspective, this implies the fraction of conduction electrons which become associated with the localized moment owing to sdexchange interaction. With no vertex correction included one obtains aG=mc 2p¯htmrD D2+2mr; (16) 4/8t=1ns t=10ns D=0.2meV D=0.3meV D=1meV 501001502002503000.0000.0010.0020.0030.004 T,KaGFigure 2. Gilbert damping, obtained from numerical integration of Eq. (8), shows almost no temperature dependence associated with thermal redistribution of conduction electrons. Dashed lines are plotted for D=1meV for t=1andt=10ns, whereas solid lines stand for D=0:2, 0:3, and 1 meV for t=100 ns. while taking account of vertex correction gives rise to a different result, aG=mc 2p¯htmrD D2+mr: (17) To provide a quantitative estimate of how large the St ˇreda contribution in the weak disorder limit is, on condition that m>jDj, we work out Q(1) mn. Using ¶gR=A(e)=¶e= [gR=A(e)]2and the fact that trace is invariant under cyclic permuattaions we conclude that only off-diagonal components m6=ncontribute. While the direct evaluation results in Q(1) xy=3mD=[2(D2+2mr)]in the clean limit. It has been demonstrated that including scattering rates dandhdoes not qualitatively change the results, leading to some smearing only52. Interestingly, within the range of applicability of theory developed in this paper, the results of both Eqs. (16) and (17) depend linearly on scattering rate, being thus in qualitative agreement with the breathing Fermi surface model. Meanwhile, the latter does not yield any connection to the microscopic parameters (see, e.g., Ref.53for more details). To provide with some quantitative estimations in our simulations we utilize the following set of parameters. Typically, experimental studies based on hyperfine field measurements equipped with DFT calculations54reveal the sdStoner interaction to be of the order of 0.2 eV , while the induced magnetization of s-derived states equals 0.002–0.05 (measured in the units of Bohr magneton, mB). Thus, the parameter of sdexchange splitting, appropriate for our model, is D0.2–1 meV . In addition, according to first-principles simulations we choose the Fermi energy m3 eV . The results of numerical integration of (8) are presented in Fig. 2 for several choices of sdexchange and scattering rates, t. The calculations reveal almost no temperature dependence in the region up to room temperature for any choice of parameters, which is associated with the fact that the dominant contribution comes from the integration in a tiny region of the Fermi energy. Fig. 2 also reveal a non-negligible dependence on the damping parameter with respect to both Dandt, which illustrates that a tailored search for materials with specific damping parameter needs to address both the sdexchange interaction as well as the scattering rate. From the theoretical perspective, the results shown in Fig. 2 correspond to the case of non-interacting electrons with no electron-phonon coupling included. Thus, the thermal effects are accounted only via temperature-induced broadening which does not show up for m>jDj. Conclusions In this paper we proposed an alternative derivation of the Gilbert damping tensor within a generalized Kubo-St ˇreda formula. We established the contribution stemming from Eq. (7) which was missing in the previous analysis within the linear response theory. In spite of being of the order of (mt) 1and, thus, negligible in the weak disorder limit developed in the paper, it should be properly worked out when dealing with more complicated systems, e.g., gapped materials such as iron garnets (certain half metallic Heusler compounds). For a model system, represented by a Rashba ferromagnet, we directly evaluated the Gilbert damping parameter and explored its behaviour associated with the temperature-dependent Fermi-Dirac distribution. In essence, the obtained results extend the previous studies within linear response theory and can be further utilized in first-principles calculations. We believe our results will be of interest in the rapidly growing fields of spintronics and magnonics. 5/8References 1.Žuti´c, I., Fabian, J. & Das Sarma, S. Spintronics: Fundamentals and applications. Rev. Mod. Phys. 76, 323 (2004). 2.Bader, S. D. & Parkin, S. S. P. Spintronics. Annu. Rev. Condens. Matter Phys. 1, 71 (2010). 3.MacDonald, A. H. & Tsoi, M. Antiferromagnetic metal spintronics. Phil. Trans. R. Soc. A 369, 3098 (2011). 4.Koopmans, B., Wagemans, W., Bloom, F. L., Bobbert, P. A., Kemerink, M. & Wohlgenannt, M. Spin in organics: a new route to spintronics. Phil. Trans. R. Soc. A 369, 3602 (2011). 5.Gomonay, E. V . & Loktev, V . M. Spintronics of antiferromagnetic systems. Low Temp. Phys. 40, 17 (2014). 6.Jungwirth, T., Marti, X., Wadley, P. & Wunderlich, J. Antiferromagnetic spintronics. Nat. Nanotechnol. 11, 231 (2016). 7.Duine, R. A., Lee, K.-J., Parkin, S. S. P. & Stiles, M. D. Synthetic antiferromagnetic spintronics. Nature Phys. 14, 217 (2018). 8.Železný, J., Wadley, P., Olejník, K., Hoffmann, A. & Ohno, H. Spin transport and spin torque in antiferromagnetic devices. Nature Phys. 14, 220 (2018). 9.Nˇemec, P., Fiebig, M., Kampfrath, T. & Kimel, A. V . Antiferromagnetic opto-spintronics. Nature Phys. 14, 229 (2018). 10.Šmejkal, L., Mokrousov, Y ., Yan, B. & MacDonald, A. H. Topological antiferromagnetic spintronics. Nature Phys. 14, 242 (2018). 11.Baltz, V ., et. al. Antiferromagnetic spintronics. Rev. Mod. Phys. 90, 015005 (2018). 12.Gilbert, T. L. A phenomenological theory of damping in ferromagnetic materials. IEEE Trans. Mag. 40, 3443 (2004). 13.Hickey, M. C. & Moodera, J. S. Origin of intrinsic Gilbert damping. Phys. Rev. Lett. 102, 137601 (2009). 14.Sharma, P., et. al. Nonvolatile ferroelectric domain wall memory. Sci. Adv. 3, 1700512 (2017). 15.Scheck, C., Cheng, L., Barsukov, I., Frait, Z. & Bailey, W. E. Low relaxation rate in epitaxial vanadium-doped ultrathin iron films. Phys. Rev. Lett. 98, 117601 (2007). 16.Woltersdorf, G., Kiessling, M., Meyer, G., Thiele, J.-U. & Back, C. H. Damping by slow relaxing rare earth impurities in Ni80Fe20.Phys. Rev. Lett. 102, 257602 (2009). 17.Zhao, Y ., et. al. Experimental investigation of temperature-dependent Gilbert damping in permalloy thin films. Sci. Rep. 6, 22890 (2016). 18.Iihama, S., et. al. Gilbert damping constants of Ta/CoFeB/MgO(Ta) thin films measured by optical detection of precessional magnetization dynamics. Phys. Rev. B 89, 174416 (2014). 19.Capua, A., Yang, S.-H., Phung, T. & Parkin, S. S. P. Determination of intrinsic damping of perpendicularly magnetized ultrathin films from time-resolved precessional magnetization measurements. Phys. Rev. B 92, 224402 (2015). 20.Heinrich, B., et. al. Ferromagnetic-resonance study of ultrathin bcc Fe(100) films grown epitaxially on fcc Ag(100) substrates. Phys. Rev. Lett. 59, 1756 (1987). 21.Platow, W., Anisimov, A. N., Dunifer, G. L., Farle, M. & Baberschke, K. Correlations between ferromagnetic-resonance linewidths and sample quality in the study of metallic ultrathin films. Phys. Rev. B 58, 5611 (1998). 22.Urban, R., Woltersdorf, G. & Heinrich, B. Gilbert damping in single and multilayer ultrathin films: Role of interfaces in nonlocal spin dynamics. Phys. Rev. Lett. 87, 217204 (2001). 23.Mizukami, S., Ando, Y . & Miyazaki, T. Effect of spin diffusion on Gilbert damping for a very thin permalloy layer in Cu/permalloy/Cu/Pt films. Phys. Rev. B 66, 104413 (2002). 24.He, P., et. al. Quadratic scaling of intrinsic Gilbert damping with spin-orbital coupling in L10FePdPt films: Experiments and ab initio calculations. Phys. Rev. Lett. 110, 077203 (2013). 25.Berger, L. Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev. B 54, 9353 (1996). 26.Šimánek, E. & Heinrich, B. Gilbert damping in magnetic multilayers. Phys. Rev. B 67, 144418 (2003). 27.Mills, D. L. Ferromagnetic resonance relaxation in ultrathin metal films: The role of the conduction electrons. Phys. Rev. B 68, 014419 (2003). 28.Tserkovnyak, Y ., Brataas, A. & Bauer, G. E. W. Enhanced Gilbert damping in thin ferromagnetic films. Phys. Rev. Lett. 88, 117601 (2002). 6/829.Tserkovnyak, Y ., Brataas, A. & Bauer, G. E. W. Spin pumping and magnetization dynamics in metallic multilayers. Phys. Rev. B 66, 224403 (2002). 30.Kamberský, V . Spin-orbital Gilbert damping in common magnetic metals. Phys. Rev. B 76, 134416 (2007). 31.Nakabayashi, N., Takeuchi, A., Hosono, K., Taguchi, K. & Tatara, G. Theory of spin relaxation torque in metallic ferromagnets. Phys. Rev. B 82, 014403 (2010). 32.Costa, A. T., Muniz, R. B., Lounis, S., Klautau, A. B. & Mills, D. L. Spin-orbit coupling and spin waves in ultrathin ferromagnets: The spin-wave Rashba effect. Phys. Rev. B 82, 014428 (2010). 33.Santos, D. L. R., Venezuela, P., Muniz, R. B. & Costa, A. T. Spin pumping and interlayer exchange coupling through palladium. Phys. Rev. B 88, 054423 (2013). 34.Gilmore, K., Idzerda, Y . U. & Stiles, M. D. Identification of the dominant precession-damping mechanism in Fe, Co, and Ni by first-principles calculations. Phys. Rev. Lett. 99, 027204 (2007). 35.Garate, I. & MacDonald, A. Gilbert damping in conducting ferromagnets. I. Kohn-Sham theory and atomic-scale inhomo- geneity. Phys. Rev. B 79, 064403 (2009). 36.Thonig, D. & Henk, J. Gilbert damping tensor within the breathing Fermi surface model: anisotropy and non-locality. New J. Phys. 16, 013032 (2014). 37.Schoen, M. A. W., et. al. Ultra-low magnetic damping of a metallic ferromagnet. Nat. Phys. 12, 839 (2016). 38.Thonig, D., Kvashnin, Y ., Eriksson, O. & Pereiro, M. Nonlocal Gilbert damping tensor within the torque-torque correlation model. Phys. Rev. Materials 2, 013801 (2018). 39.Ebert, H., Mankovsky, S., Ködderitzsch, D. & Kelly, P. J. Ab initio calculation of the Gilbert damping parameter via the linear response formalism. Phys. Rev. Lett. 107, 066603 (2011). 40.Mankovsky, S., Ködderitzsch, D., Woltersdorf, G. & Ebert, H. First-principles calculation of the Gilbert damping parameter via the linear response formalism with application to magnetic transition metals and alloys , Phys. Rev. B 87, 014430 (2013). 41.Ralph, D. C. & Stiles, M. D. Spin transfer torques , J. Magn. Magn. Mater. 320, 1190 (2008). 42.Ado, I. A., Tretiakov, O. A. & Titov, M. Microscopic theory of spin-orbit torques in two dimensions , Phys. Rev. B 95, 094401 (2017). 43.Brataas, A., Tserkovnyak, Y . & Bauer, G. E. W. Scattering theory of Gilbert damping , Phys. Rev. Lett. 101, 037207 (2008). 44.Starikov, A. A., Kelly, P. J., Brataas, A., Tserkovnyak, Y . & Bauer, G. E. W. Unified first-principles study of Gilbert damping, spin-flip diffusion, and resistivity in transition metal alloys. Phys. Rev. Lett. 105, 236601 (2010). 45.Bhattacharjee, S., Nordström, L. & Fransson, J. Atomistic spin dynamic method with both damping and moment of inertia effects included from first principles. Phys. Rev. Lett. 108, 057204 (2012). 46.Ebert, H., et. al. Calculating linear-response functions for finite temperatures on the basis of the alloy analogy model. Phys. Rev. B 91, 165132 (2015). 47.Strˇeda, P. Theory of quantised Hall conductivity in two dimensions. J. Phys. C: Solid State Phys. 15, L717 (1982). 48.Sinitsyn, N. A., MacDonald, A. H., Jungwirth, T., Dugaev, V . K. & Sinova, J. Anomalous Hall effect in a two-dimensional Dirac band: The link between the Kubo-Streda formula and the semiclassical Boltzmann equation approach. Phys. Rev. B 75, 045315 (2007). 49.Freimuth, F., Blügel, S. & Mokrousov, Y . Direct and inverse spin-orbit torques. Phys. Rev. B 92, 064415 (2015). 50.Freimuth, F., Blügel, S. & Mokrousov, Y . Chiral damping, chiral gyromagnetism, and current-induced torques in textured one-dimensional Rashba ferromagnets. Phys. Rev. B 96 , 104418 (2017). 51.Ado, I. A., Dmitriev, I. A., Ostrovsky, P. M. & Titov, M. Anomalous Hall effect in a 2D Rashba ferromagnet. Phys. Rev. Lett.117, 046601 (2016). 52.Nunner, T. S., et al. Anomalous Hall effect in a two-dimensional electron gas. Phys. Rev. B 76, 235312 (2007). 53.Eriksson, O., Bergman, A., Bergqvist, L. & Hellsvik, J. Atomistic spin dynamics: Foundations and applications (Oxford University Press, Oxford, 2017). 54.Brooks, M. S. S. & Johansson, B. Exchange integral matrices and cohesive energies of transition metal atoms. J. Phys. F: Met. Phys. 13, L197 (1983). 7/8Acknowledgements A.A.P. acknowledges the support from the Russian Science Foundation Project No. 18-72-00058. O.E. acknowledges support from eSSENCE, the Swedish Research Council (VR), the foundation for strategic research (SSF) and the Knut and Alice Wallenberg foundation (KAW). D.Y . acknowledges the support from the Russian Science Foundation Project No. 17-12-01359. Author contributions statement D.Y . conceived the idea of the paper and contributed to the theory. A.A.P. wrote the main manuscript text, performed numerical analysis and prepared figures 1-2. M.I.B. and O.E. contributed to the theory. All authors reviewed the manuscript. Additional information Competing interests The authors declare no competing interests. 8/8 | 2018-07-20 | A keen interest towards technological implications of spin-orbit driven
magnetization dynamics requests a proper theoretical description, especially in
the context of a microscopic framework, to be developed. Indeed, magnetization
dynamics is so far approached within Landau-Lifshitz-Gilbert equation which
characterizes torques on magnetization on purely phenomenological grounds.
Particularly, spin-orbit coupling does not respect spin conservation, leading
thus to angular momentum transfer to lattice and damping as a result. This
mechanism is accounted by the Gilbert damping torque which describes relaxation
of the magnetization to equilibrium. In this study we work out a microscopic
Kubo-St\v{r}eda formula for the components of the Gilbert damping tensor and
apply the elaborated formalism to a two-dimensional Rashba ferromagnet in the
weak disorder limit. We show that an exact analytical expression corresponding
to the Gilbert damping parameter manifests linear dependence on the scattering
rate and retains the constant value up to room temperature when no vibrational
degrees of freedom are present in the system. We argue that the methodology
developed in this paper can be safely applied to bilayers made of non- and
ferromagnetic metals, e.g., CoPt. | Another view on Gilbert damping in two-dimensional ferromagnets | 1807.07897v2 |
Large deviations for the Langevin equation with strong damping Sandra Cerraiy, Mark Freidlinz Department of Mathematics University of Maryland College Park, MD 20742 USA Abstract We study large deviations in the Langevin dynamics, with damping of order 1and noise of order 1, as #0. The damping coecient is assumed to be state dependent. We proceed rst with a change of time and then, we use a weak convergence approach to large deviations and their equivalent formulation in terms of the Laplace principle, to determine the good action functional. Some applications of these results to the exit problem from a domain and to the wave front propagation for a suitable class of reaction diusion equations are considered. 1 Introduction For every>0, let us consider the Langevin equation 8 >>< >>:q(t) =b(q(t)) (q(t)) _q(t) +(q(t))_B(t); q(0) =q2Rd;_q(0) =p2Rd:(1.1) HereB(t) is ar-dimensional standard Wiener process, dened on some complete stochastic basis ( ;F;fFtg;P). In what follows, we shall assume that bis Lipschitz continuous and andare bounded and continuously dierentiable, with bounded derivative. Moreover, is invertible and there exist two constants 0 < 0< 1such thata0(q)1, for all q2Rd. Equation (1.1) can be rewritten as the following system in R2d 8 >>< >>:_q(t) =p(t); q(0) =q2Rd; _p(t) =b(q(t)) (q(t)) _q(t) +(q(t))_B(t); p(0) =p2Rd; Key words : Large deviations, Laplace principle, over damped stochastic dierential equations yPartially supported by the NSF grant DMS 1407615. zPartially supported by the NSF grant DMS 1411866. 1arXiv:1503.01027v1 [math.PR] 3 Mar 2015and, due to our assumptions on the coecients, for any >0,T >0 andk1, the system above admits a unique solution z= (q;p)2Lk( ;C([0;T];R2d)), which is a Markov process. Now, if we do a change of time and dene q(t) :=q(t=),t0, we have 8 >< >:2q(t) =b(q(t)) (q(t)) _q(t) +p(q(t)) _w(t); q(0) =q2Rd;_q(0) =p 2Rd;(1.2) wherew(t) =pB(t=),t0, is another Rr-valued Wiener process, dened on the same stochastic basis ( ;F;fFtg;P). In the present paper, we are interested in studying the large deviation principle for equation (1.2), as #0. Namely, we want to prove that the family fqg>0satises a large deviation principle in C([0;T];H), with the same action functional Iand the same normalizing factor that describe the large deviation principle for the rst order equation _g(t) =b(g(t)) (g(t))+p(g(t)) (g(t))_w(t); g(0) =q2Rd: (1.3) In particular, as shown in Section 4, this implies that the asymptotic behavior of the exit time from a basin of attraction for the over damped Langevin dynamics(1.1) can be described by the quasi potential Vassociated with I, as well as the asymptotic behavior of the solutions of the degenerate parabolic and elliptic problems associated with the Langevin dynamics. Moreover, in Section 4, we will show how these results allow to prove that in reaction- diusion equations with non-linearities of KPP type, where the transport is described by the Langevin dynamics itself, the interface separating the areas where uis close to 1 and to 0, as#0, is given in terms of the action functional I, as in the classical case, when the vanishing mass approximation is considered. In [8] and [3], the system 8 >< >:q;(t) =b(q;(t)) (q;(t)) _q;(t) +p(q;(t)) _w(t); q;(0) =q2Rd; _q;(0) =p 2Rd;(1.4) for 0<;<< 1, has been studied, under the crucial assumption that the friction coecient is independent of q. It has been proven that, in this case, the so-called Kramers-Smoluchowski approxi- mation holds, that is for any xed > 0 the solution q;of system (1.4) converges in L2( ;C([0;T];Rd)), as#0, tog, the solution of the rst order equation (1.3). More- over, it has been proven that, if V(q;p) is the quasi-potential associated with the family fq;g>0, for>0 xed, then lim !0inf p2RdV(q;p) =V(q); 2whereVis the quasi-potential associated with the action-functional I. In [9], equation (1.4) with non constant friction has been considered and it has been shown that in this case the situation is considerably more delicate. Actually, the limit of q;toghas only been proven via a previous regularization of the noise, which has led to the convergence of q;to the solution ~ gof the rst order equation with Stratonovich integral. Finally, we would like to mention that in the recent paper [12], by Lyv and Roberts, an analogous problem has been studied for the stochastic damped wave equation in a bounded regular domain DRd, withd= 1;2;3, 8 >>>< >>>:@2u(t;x) @t2= u(t;x) +f(u(t;x)) @u(t;x) @t+@w(t;x) @t u(t;x) = 0; x2@D; u (0;x) =u0(x);@u(0;x) @t=v0;(x) where >0 is a small parameter, the friction coecient is constant ( = 1),w(t;x) is a smooth cylindrical Wiener process and fis a cubic non-linearity. By using the weak convergence approach, the authors show that the family fug>0satises a large deviation principle in C([0;T];L2(D)), with normalizing factor 2and the same action functional that describes the large deviation principle for the stochastic parabolic equation. As mentioned above, in the present paper we are dealing with the case of non-constant frictionand=2. Dealing with a non-constant friction coecient turns out to be important in applications, as it allows to describes new eects in reaction-diusion equations and exit problems (see section 4). Here, we will study the large deviation principle for equation(1.2) by using the approach of weak convergence (see [1] and [2]) and we will show the validity of the Laplace principle, which, together with the compactness of level sets, is equivalent to the large deviation principle. At this point, it is worth mentioning that one major diculty here is handling the integralZt 0exp Zt s(q(r))dr (q(s))dw(s); and proving that it converges to zero, as #0, inL1( ;C([0;T];Rd)). Actually, as is non- constant, the integral above cannot be interpreted as an It^ o's integral and in our estimates we cannot use It^ o's isometry. Nevertheless, due to the regularity of q(t), we can consider the integral above as a pathwise integral, and with appropriate integrations by parts, we can get the estimates required to prove the Laplace principle. 2 The problem and the method We are dealing here with the equation 8 >< >:2q(t) =b(q(t)) (q(t)) _q(t) +p(q(t)) _w(t); q(0) =q2Rd;_q(0) =p 2Rd;(2.1) 3Herew(t),t0, is ar-dimensional Brownian motion and the coecients b,andsatisfy the following conditions. Hypothesis 1. 1. The mapping b:Rd!Rdis Lipschitz-continuous and the map- ping:Rd!L(Rr;Rd)is continuously dierentiable and bounded, together with its derivative. Moreover, the matrix (q)is invertible, for any q2Rd, and 1:Rd! L(Rr;Rd)is bounded. 2. The mapping :Rd!Rbelongs toC1 b(Rd)and inf x2Rd(x) =:0>0: (2.2) In view of the conditions on the coecients ,bandassumed in Hypothesis 1, for every xed >0, equation (2.6) admits a unique solution z= (q;p)2Lk(0;T;Rd), with T >0 andk1. Now, for any predictable process utaking values in L2([0;T];Rr), we introduce the problem _gu(t) =b(gu(t)) (gu(t))+(gu(t)) (gu(t))u(t); gu(0) =q2Rd: (2.3) The existence and uniqueness of a pathwise solution guto problem (2.3) in C([0;T];Rd) is an immediate consequence of the conditions on the coecients b,andthat we have assumed in Hypothesis 1. In what follows, we shall denote by Gthe mapping G:L2([0;T];Rr)!C([0;T];Rd); u7!G(u) =gu: Moreover, for any f2C([0;T];Rd) we shall dene I(f) =1 2infZT 0ju(t)j2dt:f=G(u); u2L2([0;T];Rr) ; with the usual convention inf ;= +1. This means that I(f) =1 2ZT 0(f(s)) 1(f(s)) _f(s) b(f(s)) (f(s))2 ds; (2.4) for allf2W1;2(0;T;Rd). If we denote by gthe solution of the stochastic equation _g(t) =b(g(t)) (g(t))+p(g(t)) (g(t))_w(t); g(0) =q2Rd; (2.5) we have that Iis the large deviation action functional for the family fgg>0in the space of continuous trajectories C([0;T];Rd) (for a proof see e.g. [11]). This means that the level setsfI(f)cgare compact in C([0;T];Rd), for anyc>0, and for any closed subset FC([0;T];Rd) and any open set GC([0;T];Rd) it holds lim sup !0+logP(g2F) I(F); lim inf !0+logP(g2G) I(G); 4where, for any subset AC([0;T];Rd), we have denoted I(A) = inf f2AI(f): The main result of the present paper is to prove that in fact the family of solutions q of equation (1.2) satises a large deviation principle with the same action functional Ithat describes the large deviation principle for the family of solutions gof equation (2.5). And, due to the fact that q(t) =q(t),t0, this allows to describe the behavior of the over damped Langevin dynamics (1.1) (see Section 4 for all details). Theorem 2.1. Under Hypothesis 1, the family of probability measures fL(q)g>0, in the space of continuous paths C([0;T];Rd), satises a large deviation principle with action func- tionalI. In order to prove Theorem 2.1, we follow the weak convergence approach, as developed in [1], (see also [2]). To this purpose, we need to introduce some notations. We denote by PTthe set of predictable processes in L2( [0;T];Rr), and for any T > 0 and
>0, we dene the sets S
T= f2L2(0;T;Rd) :ZT 0jf(s)j2ds
A
T= u2PT:u2S
T;P a.s. : Next, for any predictable process utaking values in L2([0;T];Rr), we denote by qu (t) the solution of the problem 8 >< >:2qu (t) =b(qu (t)) (qu (t)) _qu (t) +p(qu (t)) _w(t) +(qu (t))u(t); qu (0) =q2Rd;_qu (0) =p 2Rd:(2.6) As well known, for any xed >0 and for any T > 0 andk1, this equation admits a unique solution qu inLk( ;C([0;T];Rd)). By proceeding as in the proof of [2, Theorem 4.3], the following result can be proven. Theorem 2.2. Letfug>0be a family of processes in S
Tthat converge in distribution, as #0, to someu2S
T, as random variables taking values in the space L2(0;T;Rd), endowed with the weak topology. If the sequencefqug>0converges in distribution to gu, as#0, in the space of contin- uous paths C([0;T];Rd), then the family fL(q)g>0satises a large deviation principle in C([0;T];Rd), with action functional I. Actually, as shown in [2], the convergence of qutoguimplies the validity of the Laplace principle with rate functional I. This means that, for any continuous mapping :C([0;T];Rd)!Rit holds lim !0 logEexp 1 (q) = inf f2C([0;T];Rd)( (f) +I(f) ): And, as the level sets of Iare compact, this is equivalent to say that fL(q)g>0satises a large deviation principle in C([0;T];Rd), with action functional I. 53 Proof of Theorem 2.1 As we have seen in the previous section, in order to prove Theorem 2.1, we have to show that iffug>0is a family of processes in S
Tthat converge in distribution, as #0, to someu2 S
T, as random variables taking values in the space L2(0;T;Rd), endowed with the weak topology, then the sequence fqug>0converges in distribution to gu, as#0, in the spaceC([0;T];Rd). In view of the Skorohod representation theorem, we can rephrase such a condition in the following way. On some probability space ( ;F;P), consider a Brownian motion wt, t0, along with the corresponding natural ltration fFtgt0. Moreover, consider a family offFtg-predictable processes fu;ug>0inL2( [0;T];Rd), taking values in S
T,P-a.s., such that the joint law of ( u;u;w), under P, coincides with the joint law of ( u;u;w ), under P, and such that lim !0u= u; P a.s. (3.1) asL2(0;T;R)-valued random variables, endowed with the weak topology. Let qube the solution of a problem analogous to (2.6), with uandwreplaced respectively by uand w. Then, we have to prove that lim !0qu=gu; P a.s. inC([0;T];Rd). In fact, we will prove more. Actually, we will show that lim !0Esup t2[0;T]jqu(t) gu(t)j= 0: (3.2) In order to prove (3.2), we will need some preliminary estimates. For any >0, we dene the process H(t) =pe A(t)Zt 0eA(s)(qu (s))dw(s); t0: (3.3) Lemma 3.1. Under Hypothesis 1, for any T > 0,k1and
> 0, there exists 0>0 such that for any u2S
Tand2(0;0] sup stEjH(t)jkck;
(T)(jqjk+jpjk+ 1)3k 2+ckk 2tk 2e k0t 2: (3.4) Moreover, we have Esup t2[0;T]jH(t)jpc
(T)(1 +jqj+jpj): (3.5) Proof. Equation (2.6) can be rewritten as the system 8 >< >:_qu (t) =pu (t); qu (0) =q 2_pu (t) =b(qu (t)) (qu (t))pu (t) +p(qu (t)) _w(t) +(qu (t))u(t); pu (0) =p : 6Thus, if for any 0 stand>0 we dene A(t;s) :=1 2Zt s(qu (r))dr; A(t) :=A(t;0); we have pu (t) =1 e A(t)p+1 2Zt 0e A(t;s)b(qu (s))ds +1 2Zt 0e A(t;s)(qu (s))u(s)ds+1 2H(t):(3.6) Integrating with respect to t, this yields qu (t) =q+1 Zt 0e A(s)pds+1 2Zt 0Zs 0e A(s;r)b(qu (r))drds +1 2Zt 0Zs 0e A(s;r)(qu (r))u(r)drds +1 2Zt 0H(s)ds:(3.7) Thanks to the Young inequality, this implies that for any t2[0;T] jqu (t)jjqj+jpj+cZt 0(1 +jqu (s)j)ds+Zt 0ju(s)jds+1 2Zt 0jH(s)jds c
(T)(jqj+jpj+ 1) +1 2Zt 0jH(s)jds+Zt 0jqu (s)jds; and from the Gronwall lemma we can conclude that jqu (t)jc
(T) (1 +jqj+jpj) +c(T)1 2Zt 0jH(s)jds: This implies that for any k1 jqu (t)jkck;
(T)(jqjk+jpjk+ 1) +ck;
(T) 2kZt 0jH(s)jkds; 2(0;1]: (3.8) Now, due to (3.6), we have jpu(t)j1 e 0t 2jpj+1 2Zt 0e 0(t s) 2(1 +jqu (s)j)ds +1 2Zt 0e 0(t s) 2ju(s)jds+1 2jH(t)j; so that, thanks to (3.8), for any 2(0;1] we get jpu (t)j1 e 0t 2jpj+c
(T)(jqj+jpj+ 1) +1 2Zt 0e 0(t s) 2ju(s)jds+c(T)1 2jH(t)j: (3.9) 7As well known, if f2C1([0;t]) andg2C([0;t]), then the Stiltjies integral Zt 0f(s)dg(s); t0; is well dened and, if g(0) = 0, the following integration by parts formula holds Zt 0f(s)dg(s) =Zt 0(g(t) g(s))h0(s)ds+g(t)h(0); t0: (3.10) Now, the mapping [0;+1)!L(Rr;Rd); s7!eA(s)(qu (s)); is dierentiable, P-a.s., so that the stochastic integral in (3.3) is in fact a pathwise integral. In particular, we can apply formula (3.10), with h(s) =eA(s)(qu (s)); g (s) =w(s); and we get H(t) =pZt 0(w(t) w(s))e A(t;s)(qu (s)) 2+0(qu (s))pu (s) ds +pw(t)e A(t)(q):(3.11) Thanks to (3.9), this yields for any 2(0;1] jH(t)jcpZt 0jw(t) w(s)je 0(t s) 2 2 1 +2jpu (s)j ds+cpjw(t)je 0t 2 c
(T)(jqj+jpj+ 1)pZt 2 0jw(t) w(t 2s)je 0sds +pc
(T)Zt 2 0jw(t) w(t 2s)je 0sjH(t 2s)jds+cpjw(t)je 0t 2; and hence, for any k1, we have jH(t)jkck;
(T)(jqjk+jpjk+ 1)k 2Zt 2 0jw(t) w(t 2s)jke 0sds +k 2ck;
(T)Zt 2 0jw(t) w(t 2s)jke 0sjH(t 2s)jkds+ckk 2jw(t)jke k0t 2: 8By taking the expectation, due to the independence of jw(t) w(t 2s)jwithjH(t 2s)j andRt 2s 0jH(r)jkdr, this implies that for any 2(0;1] EjH(t)jkck;
(T)(jqjk+jpjk+ 1)3k 2Zt 2 0sk 2e 0sds +3k 2ck;
(T)Zt 2 0sk 2e 0sEjH(t 2s)jkds+ckk 2tk 2e k0t 2 ck;
(T)(jqjk+jpjk+ 1)3k 2+ckk 2tk 2e k0t 2+3k 2ck;
(T) sup stEjH(s)jk: Therefore, if we pick 02(0;1] such that 3k 2ck;
(T)<1 2; we get (3.4). Now, let us prove (3.5). From (3.11), we have jH(t)jpcsup t2[0;T]jw(t)j 1 +Zt 0e 0(t 2) 2jpu(s)jds pcsup t2[0;T]jw(t)j 1 +Zt 0jpu(s)j2ds1 2! ; and hence Esup t2[0;T]jH(t)jpc(T) 1 + EZt 0jpu(s)j2ds1 2! : Thanks to (3.9), as a consequence of the Young inequality, we get Zt 0jpu(s)j2dsc
(T)(1 +jqj2+jpj2) +1 4c(T)Zt 0jH(s)j2ds; (3.12) so that Esup t2[0;T]jH(t)jpc
(T)(1 +jqj+jpj) +1pc(T)Zt 0EjH(s)j2ds1 2 : Therefore, (3.5) follows from (3.4). Lemma 3.2. Under Hypothesis 1, for any T >0,k1and
>0there exists 0>0such that for any u2S
Tand2(0;0)we have Esup t2[0;T]jqu (t)jkck;
(T)(jqjk+jpjk+ 1) k 2+ck;
(T)2 3k 2: (3.13) 9Proof. Estimate (3.13) follows by combining together (3.4) and (3.8). Now, we are ready to prove (3.2), that, in view of Theorem 2.2, implies Theorem 2.1. Theorem 3.3. Letfug>0be a family of predictable processes in S
Tthat converge P-a.s., as#0, to someu2S
T, with respect to the weak topology of L2(0;T;Rd). Then, we have lim !0Esup t2[0;T]jqu(t) gu(t)j= 0: (3.14) Proof. Integrating by parts in (3.7), we obtain qu(t) =q+Zt 0b(qu(s)) (qu(s))ds+Zt 0(qu(s)) (qu(s))u(s)ds+R(t); where R(t) =p Zt 0e A(s)ds 1 (qu(t))Zt 0e A(t;s)b(qu(s))ds+pZt 0(qu(s)) (qu(s))dw(s) +Zt 0Zs 0e A(s;r)b(qu(r))dr1 2(qu(s))hr(qu(s));pu(s)ids 1 (qu(t))H(t) +Zt 01 2(qu(s))H(s)hr(qu(s));pu(s)ids=:6X k=1Ik (t): This implies that qu(t) gu(t) =Zt 0b(qu(s)) (qu(s)) b(gu(s)) (gu(s)) ds+Zt 0(qu(s)) (qu(s)) (gu(s)) (gu(s)) u(s)ds +Zt 0(gu(s)) (gu(s))[u(s) u(s)]ds+R(t): (3.15) Due to the Lipschitz-continuity and the boundedness of the functions and 1=, we have that= is bounded and Lipschitz continuous. Then, as u2S
T, we obtain jqu(t) gu(t)j2 cZt 0(gu(s)) (gu(s))[u(s) u(s)]ds2 +cjR(t)j2+c(T)Zt 0jqu(s) gu(s)j2ds +c(T)Zt 0jqu(s) gu(s)j2ds Zt 0ju(s)j2ds+ sup s2[0;t]jgu(s)j2! cZt 0(gu(s)) (gu(s))[u(s) u(s)]ds2 +cjR(t)j2+c
(T)Zt 0jqu(s) gu(s)j2ds: 10By the Gronwall lemma, this allows to conclude that sup t2[0;T]jqu(t) gu(t)j c
(T) sup t2[0;T]Zt 0(gu(s)) (gu(s))[u(s) u(s)]ds+c
(T) sup t2[0;T]jR(t)j:(3.16) Now, for any >0, we dene (t) =Zt 0(gu(s)) (gu(s))[u(s) u(s)]ds: For any 0<s<t we have (t) (s) =Zt s(gu(r)) (gu(r))[u(r) u(r)]dr; so that, as uanduare both in S
T, j (t) (s)jc
p t s; > 0: As (0) = 0, this implies that the family of continuous functions is f g>0is equibounded and equicontinuous, so that, by the Ascoli-Arzel a theorem, there exists n#0 andv2 C([0;T];Rd) such that lim n!0sup t2[0;T]j n(t) v(t)j= 0;P a.s. On the other hand, as (3.1) holds, for any h2Rdwe have lim !0h (t);hi= lim !0 u u;(gu()) (gu())h L2(0;T;Rd)= 0; so that we can conclude that v= 0 and lim !0Esup t2[0;T]j (t)j= 0: Thanks to (3.16), this implies that lim sup !0Esup t2[0;T]jqu(t) gu(t)jclim sup !0Esup t2[0;T]jR(t)j; so that (3.14) follows if we show that lim !0Esup t2[0;T]jR(t)j= 0: (3.17) We have jI1 (t)j=jpj Zt 0e A(s)dscjpj 1Zt 0e 0s 2dscjpj: (3.18) 11Moreover jI2 (t)j=1 j(qu(t))jZt 0e A(t;s)b(qu(s))ds cZt 0e 0(t s) 2(1 +jqu(s)j)dsc2 1 + sup t2[0;T]jqu(t)j! : Thanks to (3.13), this implies Esup t2[0;T]jI2 (t)jc
(T)(jpj+jqj+ 1)3 2; 2(0;1]: (3.19) Next Esup t2[0;T]jI3 (t)j=pEsup t2[0;T]Zt 0(qu(s)) (qu(s))dw(s)c(T)p: (3.20) Concerning I4(t), we have jI4 (t)j=Zt 0Zs 0e A(s;r)b(qu(r))dr1 2(qu(s))hr(qu(s));pu(s)ids 2c 1 + sup t2[0;T]jqu(t)j!Zt 0jpu(s)jds; so that, due to (3.13) we obtain Esup t2[0;T]jI4 (t)j2c
(T)(jqj+jpj+ 1) 1 2 EZt 0jpu(s)j2ds1 2 : As a consequence of (3.4) and (3.12), this yields Esup t2[0;T]jI4 (t)jc
(T)(jqj2+jpj2+ 1); 2(0;0]: (3.21) Concerning I5 (t), according to (3.5) we have Esup t2[0;T]jI5 (t)jcEsup t2[0;T]jH(t)jpc
(T)(1 +jqj+jpj): (3.22) Finally, it remains to estimate I6 (t). We have jI6 (t)j=Zt 01 2(qu(s))H(s)hr(qu(s));pu(s)idscZt 0jH(s)jjpu(s)jds; so that Esup t2[0;T]jI6 (t)jcZT 0EjH(s)j2dsZT 0Ejpu(s)j2ds1 2 : 12By using (3.12), this gives Esup t2[0;T]jI6 (t)jc
(T)(1 +jqj+jpj)ZT 0EjH(s)j2ds1 2 +1 2ZT 0EjH(s)j2ds; so that, from (3.4) we get Esup t2[0;T]jI6 (t)jc
(T)(1 +jqj+jpj); 2(0;0]: This, together with (3.18), (3.19), (3.20), (3.21) and (3.22), implies (3.17) and (3.14) follows. 4 Some applications and remarks LetGbe a bounded domain in Rd, with a smooth boundary @G. We consider here the exit problem for the process q(t) dened as the solution of equation (1.1). For every >0 we dene := minft0 :q(t)=2Gg; := minft0 :q(t)=2Gg; whereq(t) =q(t=) is the solution of equation (2.6). It is clear that =1 ; q() =q(): In what follows, we shall assume that the dynamical system _q(t) =b(q(t)); t0; (4.1) satises the following conditions. Hypothesis 2. The pointO2Gis asymptotically stable for the dynamical system (4.1) and for any initial condition q2Rd lim t!1q(t) =O: Moreover, we have hb(q);(q)i>0; q2@G; where(q)is the inward normal vector at q2@G. Now, we introduce the quasi-potential associated with the action functional Idened in (2.4) V(q) = infn I(f); f2C([0;T];Rd); f(0) =O; f (T) =q; T > 0o =1 2inf(ZT 0(f(s)) 1(f(s)) _f(s) b(f(s)) (f(s))2 ds; f (0) =O; f (T) =q; T > 0) : It is easy to check that, under our assumptions on (q), the quasi-potential Vcoincides with 1 2infZT 0 1(f(s)) _f(s) (f(s))b(f(s))2 ds; f (0) =O; f (T) =q; T > 0 :(4.2) 13Theorem 4.1. Under Hypotheses 1 and 2, for each q2fq2G:V(q)V0gandp2Rd, we have lim !0logE(q;p)= lim !0logE(q;p)=V0; (4.3) and lim !0log= lim !0log=V0;in probability ; (4.4) where V0:= min q2@GV(q): Moreover, if the minimum of Von@Gis achieved at a unique point q?2@G, then lim !0q() = lim !0q() =q?: (4.5) Proof. First, note that q(t) is the rst component of the 2 d-dimensional Markov process z(t) = (q(t);p(t)). Because of the structure of the p-component of the drift of this process and our assumptions on the vector eld b, starting from ( q;p)2R2d, the trajectory of z(t) spends most of the time in a small neighborhood of the point q=Oandp= 0, with probability close to 1, as 0 < << 1. From time to time, the process z(t) deviates from this point and, as proven in Theorem 2.1, the deviations of q(t) are governed by the large deviation principle with action functional I, dened in (2.4). This allows to prove the validity of (4.3), (4.4) and (4.5) in the same way as Theorems 4.41, 4.42 and 4.2.1 from [11] are proven. We omit the details. As an immediate consequence of (4.2) and [11, Theorem 4.3.1], we have the following result. Theorem 4.2. Assumea(q) :=(q)?(q) =Iand(q)b(q) = rU(q) +l(q), for any q2Rd, for some smooth function U:Rd!Rhaving a unique critical point (a minimum) atO2Rdand such that hrU(q);l(q)i= 0; q2Rd: Then V(q) = 2U(q); q2Rd: From Theorems 4.1 and 4.2, it is possible to get a number of results concerning the asymptotic behavior, as #0, of the solutions of the degenerate parabolic and the elliptic problems associated with the dierential operator Ldened by Lu(q;p) =1 2dX i;j=1ai;j(q)@2u @pi@pj(q;p) + b(q) 1 (q)p rpu(q;p) +prqu(q;p): Assume now that the dynamical system (4.1) has several asymptotically stable attrac- tors. Assume, for the sake of brevity, that all attractors are just stable equilibriums O1, O2,. . . ,Ol. Denote byEthe set of separatrices separating the basins of these attractors, and assume the setEto have dimension strictly less than d. Moreover, let each trajectory q(t), 14starting at q02RdnE, be attracted to one of the stable equilibriums Oi,i= 1;:::;l , as t!1 . Finally, assume that the projection of b(q) on the radius connecting the origin in Rdand the point q2Rdis directed to the origin and its length is bounded from below by some uniform constant >0 (this condition provides the positive recurrence of the process z(t) = (q(t);p(t)),t0). In what follows, we shall denote V(q1;q2) =1 2infZT 0(f(s)) 1(f(s)) _f(s) b(f(s)) (f(s))2 ds; f (0) =q1; f(T) =q2; T > 0 and Vij=V(Oi;Oj); i;j2f1;:::;lg: In a generic case, the behavior of the process ( q(t);p(t)), on time intervals of order exp( 1), > 0 and 0< << 1, can be described by a hierarchy of cycles as in [11] and [6]. The cycles are dened by the numbers Vij. For (almost) each initial point qand a time scale, these numbers dene also the metastable state Oi?,i?=i?(q;), whereq(t) spends most of the time during the time interval [0 ;exp( 1)]. Slow changes of the eld b(q) and/or of the damping coecient (q) can lead to stochastic resonance (compare with [7]). Consider next the reaction diusion equation in Rd 8 >< >:@u @t(t;q) =Lu(t;q) +c(q;u(t;q))u(t;q); u(0;q) =g(q); q2Rd; t> 0:(4.6) HereLis a linear second order uniformly elliptic operator, with regular enough coecients. Letq(t) be the diusion process in Rdassociated with the operator L. The Feynman-Kac formula says that ucan be seen as the solution of the problem u(t;q) =Eqg(q(t)) expZt 0c(q(s);u(t s;q(s))ds: (4.7) Reaction-diusion equations describe the interaction between particle transport dened byq(t) and reaction which consists of multiplication (if c(q;u)>0) and annihilation (if c(q;u)<0) of particles. In classical reaction-diusion equations, the Langevin dynamics which describes a diusion with inertia is replaced by its vanishing mass approximation. If the transport is described by the Langevin dynamics itself, equation (4.6) should be replaced by an equation in R2d. Assuming that the drift is equal to zero ( b(q) = 0), and the damping is of order 1, as#0, this equation has the form8 >>>>>>>>< >>>>>>>>:@u @t(t;q;p ) =1 2dX i;j=1ai;j(q)@2u @pi@pj(q;p) 1 (q)prpu(q;p) +prqu(q;p) +c(q;u(t;q;p ))u(t;q;p ); t> 0;(q;p)2R2d; u(0;q;p) =g(q)0;(q;p)2R2d:(4.8) 15Now, we dene R(t;q) = supZt 0c(f(s);0)ds It(f) :f(0) =q; f(t)2G0 ; where It(f) =1 2Zt 02(f(s))a 1(f(s))_f(s)_f(s)ds; andG0= suppfg(q); q2Rdg. Denition 4.3. 1. We say that Condition (N) is satised if R(t;x)can be characterized, for anyt>0andx2t=fq2Rd; R(t;q) = 0g, as supZt 0c(f(s);0)ds It(f); f(0) =q; f(t)2G0; R(t s;f(s))0;0st : 2. We say that the non-linear term f(q;u) =c(q;u)uin equation (4.8) is of KPP (Kolmogorov-Petrovskii-Piskunov) type if c(q;u)is Lipschitz-continuous, c(q;0) c(q;u)>0, for any 0<u< 1,c(q;1) = 0 andc(q;u)<0, for anyu>1. Theorem 4.4. Let the non-linear term in (4.8) be of KPP type. Assume that Condition (N) is satised and assume that the closure of G0= suppfg(q); q2Rdgcoincides with the closure of the interior of G0. Then, lim !0u(t=;q;p ) = 0;ifR(t;q)<0; (4.9) and lim !0u(t=;q;p ) = 1;ifR(t;q)>0; (4.10) so that equation R(t;q) = 0 inR2ddenes the interface separating the area where u, the solution of (4.8) , is close to 1and to 0, as#0. Proof. If we dene u(t;q;p ) =u(t=;q;p ), the analog of (4.7) yields u(t;q;p ) =E(q;p)g(q(t)) exp1 Zt 0c(q(s);u(t s;q(s);p(s))ds ; (4.11) wherez(t) = (q(s);p(s)) is the solution to equation (2.6). By taking into account our assumptions on c(q;u), we derive from (4.11) u(t;q;p )E(q;p)g(q(t)) exp1 Zt 0c(q(s);0)ds : Theorem 2.1 and the Laplace formula imply that the right hand side of the above inequality is logarithmically equivalent , as #0, to exp 1 R(t;q) and this implies (4.9). In order to prove (4.10), rst of all one should check that if R(t;q) = 0, then for each >0 u(t;q;p )exp 1 ; (4.12) 16when>0 is small enough. This follows from (4.11) and Condition (N), if one takes into account the continuity of c(q;u). The strong Markov property of the process ( q(t);p(t)) and bound (4.12) imply (4.10) (compare with [5]). Consider, as an example, the case c(q;0) =c= const. Then R(t;q) =ct inffIt(f); f(0) =q; f(t)2G0g: The inmum in the equality above coincides with 1 2t2(q;G 0); (see, for instance, [5] for a proof), where (q1;q2),q1;q22Rd, is the distance in the Riemaniann metric ds=(q)vuutdX i;j=1ai;j(q)dqidqj: This implies that the interface moves according to the Huygens principle with the constant speedp 2c, if calculated in the Riemannian metric ds. If(q) = 0 in a domain G1Rd, the points of G1should be identied. The Riemaniann metric in Rdinduces now, in a natural way, a new metric ~ in this space with identied points. The motion of the interface, in this case, can be described by the Huygens principle with constant velocityp 2cin the metric ~ . Ifc(q;0) is not constant, the motion of the interface, in general, cannot be described by a Huygens principle. Actually, the motion can have jumps and other specic features (compare with [5]). Finally, if the Condition (N) is not satised, the function R(t;q) should be replaced by another one. Dene ~R(t;q) = sup min 0atZa 0c(f(s);0)ds Ia(f) :f(0) =q; f(t)2G0 : The function ~R(t;q) is Lipschitz continuous and non-positive and if Condition (N) is satis- ed, then ~R(t;q) = minfR(t;q);0g: By proceeding as in [6], it is possible to prove that lim !0u(t=;q;p ) = 0;ifR(t;q)<0; and lim !0u(t=;q;p ) = 1; if (t;q) is in the interior of the set f(t;q) :t>0; q2Rd;~R(t;q) = 0g. Finally, we would like to mention a few generalizations. 171. The arguments that we we have used in the proof of Theorem 2.1, can be used to prove the same result for the equation 8 >>< >>:q(t) =b(q(t)) (q(t)) _q(t) +1 (q(t))_B(t); q(0) =q2Rd;_q(0) =p2Rd; for any <1=2. As a matter of fact, with the very same method we can show that also in this case the family fqg>0satises a large deviation principle in C([0;T];Rd) with action functional Iand with normalizing factor 1 2. 2. The damping can be assumed to be anisotropic. This means that the coecient (q) can be replaced by a matrix (q), with all eigenvalues having negative real part. 3. Systems with strong non-linear damping can be considered. Namely, let ( q;p) be the time-inhomogeneous Markov process corresponding to the following initial-boundary value problem for a degenerate quasi-linear equation on a bounded regular domain GRd 8 >>>>>>>>>< >>>>>>>>>:@u(t;p;q ) @t=1 2dX i;j=1ai;j(q)@2u(t;q;p ) @pi@pj+b(q)rpu(t;q;p ) (q;u(t;q;p )) prpu(t;q;p ) +prqu(t;q;p ): u(0;q;p) =g(q); u(t;q;p )jq2@G= (q); Existence and uniqueness of such degenerate problem, under some mild conditions, follows from [4, Chapter 5]. The non-linearity of the damping leads to some pecu- larities in the exit problem and in metastability. In particular, in the generic case, metastable distributions can be distributions among several asymptotic attractors and the limiting exit distributions may have a density (see [10]). References [1] P. Dupuis, R. Ellis, A weak convergence approach to the theory of large deviations , Wiley Series in Probability and Statistics, John Wiley and Sons, Inc., New York, 1997. [2] M. Bou e, P. Dupuis, A variational representation for certain functionals of Brownian motion , Annals of Probability 26 (1998), no. 4, 1641{1659. [3] Z. Chen, M.I. Freidlin, Smoluchowski-Kramers approximation and exit problems Stochastics and Dynamycs 5 (2005), pp. 569{585. [4] M.I. Freidlin, Functional integration and partial differential equations , Annals of Mathematics Studies, 109, Princeton University Press, 1985. 18[5] M.I. Freidlin, Limit theorems for large deviations and reaction-diusion equations , An- nals of Probability 13 (1985), pp. 639{675. [6] M.I. Freidlin, Coupled reaction-diusion equations , Annals of Probability 19 (1991), pp. 29{57. [7] M.I. Freidlin, Quasi-deterministic approximation, metastability and stochastic reso- nance , Physica D 137 (2000), pp. 333{352. [8] M.I. Freidlin, Some remarks on the Smoluchowski-Kramers approximation , Journal of Statistical Physics 117, pp. 617{634, 2004. [9] M.I. Freidlin, W. Hu, Smoluchowski-Kramers approximation in the case of variable friction , Journal of Mathematical Sciences, 179 (2011), pp. 184{207. [10] M.I. Freidlin, L. Koralov, Nonlinear stochastic perturbations of dynamical systems and quasi-linear parabolic PDE's with a small parameter , Probability Theory andRelated Fields 147 (2010), pp. 273{301. [11] M.I. Freidlin, A.D. Wentzell, Random perturbations of dynamical systems , Third Edition, Springer, Heidelberg, 2012. [12] Y. Lyv, A.J. Roberts, Large deviation principle for singularly perturbed stochastic damped wave equations , Stochastic Analysis and Applications, 32 (2014), pp. 50-60. 19 | 2015-03-03 | We study large deviations in the Langevin dynamics, with damping of order
$\e^{-1}$ and noise of order $1$, as $\e\downarrow 0$. The damping coefficient
is assumed to be state dependent. We proceed first with a change of time and
then, we use a weak convergence approach to large deviations and their
equivalent formulation in terms of the Laplace principle, to determine the good
action functional.
Some applications of these results to the exit problem from a domain and to
the wave front propagation for a suitable class of reaction diffusion equations
are considered. | Large Deviations for the Langevin equation with strong damping | 1503.01027v1 |
Inertia, diffusion and dynamics of a driven skyrmion Christoph Sch ¨utte,1Junichi Iwasaki,2Achim Rosch,1and Naoto Nagaosa2, 3, 1Institut f ¨ur Theoretische Physik, Universit ¨at zu K ¨oln, D-50937 Cologne, Germany 2Department of Applied Physics, University of Tokyo, 7-3-1, Hongo, Bunkyo-ku, Tokyo 113-8656, Japan 3RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan (Dated: January 5, 2015) Skyrmions recently discovered in chiral magnets are a promising candidate for magnetic storage devices because of their topological stability, small size ( 3 100nm), and ultra-low threshold current density ( 106A/m2) to drive their motion. However, the time-dependent dynamics has hitherto been largely unexplored. Here we show, by combining the numerical solution of the Landau-Lifshitz-Gilbert equation and the analysis of a generalized Thiele’s equation, that inertial effects are almost completely absent in skyrmion dynamics driven by a time-dependent current. In contrast, the response to time-dependent magnetic forces and thermal fluctuations depends strongly on frequency and is described by a large effective mass and a (anti-) damping depending on the acceleration of the skyrmion. Thermal diffusion is strongly suppressed by the cyclotron motion and is proportional to the Gilbert damping coefficient . This indicates that the skyrmion position is stable, and its motion responds to the time-dependent current without delay or retardation even if it is fast. These findings demonstrate the advantages of skyrmions as information carriers. PACS numbers: 73.43.Cd,72.25.-b,72.80.-r I. INTRODUCTION Mass is a fundamental quantity of a particle determining its mechanical inertia and therefore the speed of response to ex- ternal forces. Furthermore, it controls the strength of quantum and thermal fluctuations. For a fast response one usually needs small masses and small friction coefficients which in turn lead to large fluctuations and a rapid diffusion. Therefore, usually small fluctuations and a quick reaction to external forces are not concomitant. However a “particle” is not a trivial object in modern physics, it can be a complex of energy and mo- mentum, embedded in a fluctuating environment. Therefore, its dynamics can be different from that of a Newtonian parti- cle. This is the case in magnets, where such a “particle” can be formed by a magnetic texture1,2. A skyrmion3,4is a rep- resentative example: a swirling spin texture characterized by a topological index counting the number of times a sphere is wrapped in spin space. This topological index remains un- changed provided spin configurations vary slowly, i.e., dis- continuous spin configurations are forbidden on an atomic scale due to high energy costs. Therefore, the skyrmion is topologically protected and has a long lifetime, in sharp con- trast to e.g. spin wave excitations which can rapidly decay. Skyrmions have attracted recent intensive interest because of their nano-metric size and high mobility5–14. Especially, the current densities needed to drive their motion ( 106A/m2) are ultra small compared to those used to manipulate domain walls in ferromagnets ( 1011 12A/m2)15–19. The motion of the skyrmion in a two dimensional film can be described by a modified version of Newton’s equation. For sufficiently slowly varying and not too strong forces, a sym- metry analysis suggests the following form of the equations of motion, G_R+D_R+mR+ R=Fc+Fg+Fth:(1) Here we assumed translational and rotational invariance of the linearized equations of motion. The ‘gyrocoupling’ G=G^e?is an effective magnetic field oriented perpendicular to the plane,is the (dimensionless) Gilbert damping of a sin- gle spin,Ddescribes the friction of the skyrmion, mits mass andRits centre coordinate. parametrizes a peculiar type of damping proportional to the acceleration of the particle. We name this term ‘gyrodamping’, since it describes the damping of a particle on a cyclotron orbit (an orbit with R/G_R), which can be stronger ( parallel to G) or weaker (antipar- allel to G) than that for linear motion. Our main goal will be to describe the influence of forces on the skyrmion arising from electric currents ( Fc), magnetic field gradients (Fg) and thermal fluctuations (Fth). By analyzing the motion of a rigid magnetic structure M(r;t) =M0(r R(t))forstatic forces, one can obtain analytic formulas for G;D;FcandFgusing the approach of Thiele19–22,24. In Ref. [25], an approximate value for the mass of a skyrmion was obtained by simulating the motion of a skyrmion in a nanodisc and by estimating contributions to the mass from internal excitations of the skyrmion. For rapidly changing forces, needed for the manipulation of skyrmions in spintronic devices, Eq. (1) is however not suffi- cient. A generalized version of Eq. (1) valid for weak but also arbitrarily time-dependent forces can be written as G 1(!)V(!) =Fc(!) +Fg(!) +Fth(!) (2) =Sc(!)vs(!) +Sg(!)rBz(!) +Fth(!) HereV(!) =R ei!t_R(t)dtis the Fourier transform of the velocity of the skyrmion, vs(!)is the (spin-) drift velocity of the conduction electrons, directly proportional to the cur- rent,rBz(!)describes a magnetic field gradient in frequency space. The role of the random thermal forces, Fth(!), is spe- cial as their dynamics is directly linked via the fluctuation- dissipation theorem to the left-hand side of the equation, see below. The 22matrix G 1(!)describes the dynam- ics of the skyrmion; its small- !expansion defines the terms written on the left-hand side of Eq. (1). One can expectarXiv:1501.00444v1 [cond-mat.str-el] 2 Jan 20152 FIG. 1: When a skyrmion is driven by a time dependent external force, it becomes distorted and the spins precess resulting in a de- layed response and a large effective mass. In contrast, when the skyrmion motion is driven by an electric current, the skyrmion ap- proximately flows with the current with little distortion and preces- sion. Therefore skyrmions respond quickly to rapid changes of the electric current. strongly frequency-dependent dynamics for the skyrmion be- cause the external forces in combination with the motion of the skyrmion can induce a precession of the spin and also ex- cite spinwaves in the surrounding ferromagnet, see Fig. 1. We will, however, show that the frequency dependence of the right-hand side of the Eq. (2) is at least as important: not only the motion of the skyrmion but also the external forces excite internal modes. Depending on the frequency range, there is an effective screening or antiscreening of the forces described by the matrices Sc(!)andSg(!). Especially for the current- driven motion, there will be for all frequencies an almost exact cancellation of terms from G 1(!)andSc(!). As a result the skyrmion will follow almost instantaneously any change of the current despite its large mass. In this paper, we study the dynamics of a driven skyrmion by solving numerically the stochastic Landau-Lifshitz-Gilbert (LLG) equation. Our strategy will be to determine the param- eters of Eq. (2) such that this equation reproduces the results of the LLG equation. Section II introduces the model and outlines the numerical implementation. Three driving mecha- nisms are considered: section III studies the diffusive motion of the skyrmion due to thermal noise, section IV the skyrmion motion due to time-dependent magnetic field gradient and sec- tion V the current-driven dynamics. We conclude with a sum- mary and discussion of the results in Sec. VI. II. MODEL Our study is based on a numerical analysis of the stochastic Landau-Lifshitz-Gilbert (sLLG) equations27defined by dMr dt=
Mr[Be+b
(t)]
MMr(Mr[Be+b
(t)]):(3) Here
is the gyromagnetic moment and the Gilbert damp- ing;Be= H[M] Mris an effective magnetic field created by the surrounding magnetic moments and b
(t)a fluctuating,stochastic field creating random torques on the magnetic mo- ments to model the effects of thermal fluctuations, see below. The Hamiltonian H[M]is given by H[M] = JX rMr Mr+aex+Mr+aey X r MrMr+aexex+MrMr+aeyey BX rMr (4) We useJ= 1 ,
= 1 ,jMrj= 1 ,= 0:18Jfor the strength of the Dzyaloshinskii-Moriya interaction and B= (0;0;0:0278J)for all plots giving rise to a skyrmion with a radius of about 15lattice sites, see Appendix A. For this pa- rameter set, the ground state is ferromagnetic, thus the single skyrmion is a topologically protected, metastable excitation. Typically we simulate 100100spins for the analysis of dif- fusive and current driven motion and 200200spins for the force-driven motion. For these parameters lattice effects are negligible, see appendix B. Typical microscopic parameters used, areJ= 1meV (this yields Tc10K) which we use to estimate typical time scales for the skyrmion motion. Following Ref. 27, we assume that the field bfl r(t)is gen- erated from a Gaussian stochastic process with the following statistical properties bfl r;i(t) = 0 bfl r;i(t)bfl r0;j(s) = 2kBT
Mijrr0(t s) (5) whereiandjare cartesian components and h:::idenotes an average taken over different realizations of the fluctuating field. The Gaussian property of the process stems from the in- teraction of Mrwith a large number of microscopic degrees of freedom (central limit theorem) which are also responsi- ble for the damping described by , reflecting the fluctuation- dissipation theorem. The delta-correlation in time and space in Eq. (5) expresses that the autocorrelation time and length of bfl r(t)is much shorter than the response time and length scale of the magnetic system. For a numerical implementation of Eq. (3) we follow Ref. 27 and use Heun’s scheme for the numerical integration which converges quadratically to the solution of the general system of stochastic differential equations (when interpreted in terms of the Stratonovich calculus). For static driving forces, one can calculate the drift veloc- ity_Rfollowing Thiele20. Starting from the Landau-Lifshitz Gilbert equations, Eq. (3), we project onto the translational mode by multiplying Eq. (3) with @iMrand integrating over space21–23. G=~M0Z dr n(@xn@yn) D=~M0Z dr(@xn@xn+@yn@yn)=2 Fc=Gvs+Dvs; Fg=MsrB; M s=M0Z dr(1 nz) (6)3 where nis the direction of the magnetization, M0the lo- cal spin density, vsthe (spin-) drift velocity of the conduc- tion electrons proportional to the electric current, and Ms is the change of the magnetization induced by a skyrmion in a ferromagnetic background. The ’gyrocoupling vector’ G= (0;0;G)TwithG=~M04is given by the winding number of the skyrmion, independent of microscopic details. III. THERMAL DIFFUSION Random forces arising from thermal fluctuations play a de- cisive role in controlling the diffusion of particles and there- fore also the trajectories R(t)of a skyrmion. To obtain R(t) and corresponding correlation functions we used numerical simulations based on the stochastic Landau-Lifshitz-Gilbert equation27. These micromagnetic equations describe the dy- namics of coupled spins including the effects of damping and thermal fluctuations. Initially, a skyrmion spin-texture is embedded in a ferromagnetic background. By monitoring the change of the magnetization, we track the center of the skyrmion R(t), see appendix A for details. Our goal is to use this data to determine the matrix G 1(!) and the randomly fluctuating thermal forces, Fth(!), which together fix the equation of motion, Eq. (2), in the presence of thermal fluctuations ( rBz=vs= 0). One might worry that this problem does not have a unique solution as both the left-hand and the right-hand side of Eq. (2) are not known a priori. Here one can, however, make use of the fact that Kubo’s fluctuation-dissipation theorem26constraints the ther- mal forces on the skyrmion described by Fthin Eq. (2) by linking them directly to the dissipative contributions of G 1. On averagehFth= 0i, but its autocorrelation is proportional to the temperature and friction coefficients. In general it is given by hFi th(!)Fj th(!0)i=kBT[G 1 ij(!) +G 1 ji( !)]2(!+!0): (7) For small ! one obtainshFx th(!)Fx th(!0)i = 4kBTD(!+!0)while off-diagonal correla- tions arise from the gyrodamping hFx th(!)Fy th(!0)i= 4i!kBT (!+!0). Using Eq. (7) and demand- ing furthermore that the solution of Eq. (2) reproduces the correlation function h_Ri(t)_Rj(t0)i(or, equivalently, h(Ri(t) Rj(t0))2i) obtained from the micromagnetic simulations, leads to the condition26 Gij(!) =1 kBTZ1 0(t t0)h_Ri(t)_Rj(t0)i (8) ei!(t t0)d(t t0): We therefore determine first in the presence of thermal fluc- tuations (rBz=vs= 0) from simulations of the stochastic LLG equation (3) the correlation functions of the velocities and use those to determine Gij(!)using Eq. (8). After a sim- ple matrix inversion, this fixes the left-hand side of the equa- tion of motion, Eq. (2), and therefore contains all information 0 5 10 15 20 25t ωp00.511.522.5 <ΔR2>α=0.01 α=0.05 α=0.1 α=0.15 α=0.2FIG. 2: Time dependence of the correlation function (Ri(t0+t) Ri(t0))2 forT= 0:1Jand different values of the Gilbert damping (!p=B= 0:0278Jis the frequency for cyclotron motion). on the (frequency-dependent) effective mass, gyrocoupling, damping and gyrodamping of the skyrmion. Furthermore, the corresponding spectrum of thermal fluctuations is given by Eq. (7). Fig. 2 showsh(R)2it=h(Rx(t0+t) Rx(t0))2i. As expected, the motion of the skyrmion is diffusive: the mean squared displacement grows for long times linearly in time h(R)2it= 2Dt, whereDis the diffusion constant. Usu- ally the diffusion constant of a particle grows when the fric- tion is lowered26. For the skyrmion the situation is opposite: the diffusion constant becomes small for the small friction, i.e., small Gilbert damping . This surprising observation has its origin in the gyrocoupling G: in the absence of friction the skyrmion would be localized on a cyclotron orbit. From Eq. (1), we obtain D=kBTD G2+ (D)2(9) The diffusion is strongly suppressed by G. As in most materi- alsis much smaller than unity while DG , the skyrmion motion is characterized both by a small diffusion constant and a small friction. Such a suppressed dynamics has also been shown to be important for the dynamics of magnetic vortices28. For typical parameters relevant for materials like MnSi we estimate that it takes 10 6sto10 5sfor a skyrmion to diffusive over an average length of one skyrmion diameter. To analyze the dynamics on shorter time scales we show in Fig. 3 four real functions parametrizing G 1(!): a frequency- dependent mass m(!), gyrocouplingG(!), gyrodamping (!)and dissipation strength D(!)with G 1(!) = D(!) i!m(!) G(!) +i! (!) G(!) i! (!)D(!) i!m(!) For!!0one obtains the parameters of Eq. (1). All pa- rameters depend only weakly on temperature, Gandmare ap- proximately independent of , while the friction coefficients4 0 1 2 3 4 5 ω / ωp04812 α thermal diffusion 0 1 2 3 4 5 ω / ωp00.51-G / 4 π 0 1 2 3 4 5 ω / ωp0100200 α Γ0 1 2 3 4 5 ω / ωp050100 m current driven motion force driven motion FIG. 3: Dissipative tensor D, massm, gyrocouplingGand gyro- damping as functions of the frequency !for the diffusive motion atT= 0:1(solid lines). They differ strongly from the “apparent” dynamical coefficients (see text) obtained for the force driven (red dashed line) and current driven motion (green dot-dashed line). We use= 0:2,= 0:1. The error bars reflect estimates of systematic errors arising mainly from discretization effects, see appendix B. 00.05 0.1 0.15 0.2α01234 αT=0.15 T=0.2 00.05 0.1 0.15 0.2α00.51-G / 4 π 00.05 0.1 0.15 0.2α010203040 α Γ00.05 0.1 0.15 0.2α0255075100 mT=0.05 T=0.1 FIG. 4: Dissipative strength D, massm, gyrocouplingGand gy- rodamping as functions of the Gilbert damping for different temperatures T. Dand are linear in , see Fig. 4. In the limit T!0, G(!!0)takes the value 4, fixed by the topology of the skyrmion15,20. Both the gyrodamping and and the effective mass m have huge numerical values. A simple scaling analysis of the Landau-Lifshitz-Gilbert equation reveals that both the gyro- couplingGandDare independent of the size of the skyrmion, while andmare proportional to the area of the skyrmion, and frequencies scale with the inverse area, see appendix B. For the chosen parameters (the field dependence is dis-cussed in the appendix B), we find m0:3N
ipm0and 0:7N
ipm0, wherem0=~2 Ja2is the mass of a sin- gle flipped spin in a ferromagnet ( 1in our units) and we have estimated the number of flipped spins, N
ip, from the total magnetization of the skyrmion relative to the ferromagnetic background. As expected the mass of skyrmions grows with the area (consistent with an estimate29formobtained from the magnon spectrum of skyrmion crystals), the observation that the damping rate Dis independent of the size of skyrmions is counter-intuitive. The reason is that larger skyrmions have smoother magnetic configurations, which give rise to less damping. For realistic system parameters J= 1meV (which yields a paramagnetic transition temperature TC10K, but there are also materials, i.e. FeGe, where the skyrmion lattice phase is stabilised near room-temperature16) anda= 5 ˚A and a skyrmion radius of 200 ˚A one finds a typical mass scale of 10 25kg. The sign of the gyrodamping is opposite to that of the gyrocouplingG. This implies that describes not damp- ing but rather antidamping: there is less friction for cyclotron motion of the skyrmion than for the linear motion. The nu- merical value for the antidamping turns out to be so large thatDm+ G<0. This has the profound consequence that the simplified equation of motion shown in Eq. (1) cannot be used: it would wrongly predict that some oscillations of the skyrmion are not damped, but grow exponentially in time due to the strong antidamping. This is, however, a pure artifact of ignoring the frequency dependence of G 1(!), and such oscillations do not grow. Fig. 3 shows that the dynamics of the skyrmion has a strong frequency dependence. We identify the origin of this fre- quency dependence with a coupling of the skyrmion coordi- nate to pairs of magnon excitations as discussed in Ref. 31. Magnon emission sets in for ! > 2!pwhere!p=Bis the precession frequency of spins in the ferromagnet (in the pres- ence of a bound state with frequency !b, the onset frequency is!p+!b, Ref. 31). This new damping channel is most ef- ficient when the emitted spin waves have a wavelength of the order of the skyrmion radius. As a test for this mechanism, we have checked that only this high-frequency damping survives for !0. In Fig. 5 we show the frequency dependent damping D(!)for various bare damping coefficients . For small!it is proportional to as predicted by the Thiele equation. For !>2!p, however, an extra dampling mechanism sets in: the skyrmion motion can be damped by the emission of pairs of spin waves. This mechanism is approximately independent of and survives in the!0limit. This leads necessarily to a pronounced frequency dependence of the damping and therefore to the ef- fective mass m(!)which is related by the Kramers-Kronig relationm(!) =1 !R1 1D(!0) !0 !d!0 toD(!). Note also that the largeindependent mass m(!!0)is directly related to theindependent damping mechanism for large !. Also the frequency dependence of m(!)andG(!)can be traced back to the same mechanism as these quantities can be computed fromD(!)and (!)using Kramers-Kronig relations. For large frequencies, the effective mass practically vanishes and5 0 1 2 3 4 5 ω/ωp05101520αD(ω)α=0.05 α=0.1 α=0.2 FIG. 5: Effective damping, D(!)for= 0:2,0:1and0:05. 0 1 2 3 4 5 ω / ωp-400-2000200400Mz tot600Re/Im SgRe Sg11 Re Sg21 Im Sg11 Im Sg21 FIG. 6: Dynamical coupling coefficients for the force driven motion (= 0:2). In the static limit everything but the real part of the diago- nal vanishes. R eS11 g(!)however approaches the total magnetization Mz totas expected. The error bars reflect estimates of systematic er- rors, see appendix B. the ‘gyrocoupling’ Gdrops by a factor of a half. IV . FORCE-DRIVEN MOTION Next, we study the effects of an oscillating magnetic field gradient rBz(t)in the absence of thermal fluctuations. As the skyrmion has a large magnetic moment Mz totrelative to the ferromagnetic background, the field gradient leads to a force acting on the skyrmion. In the static limit, the force is exactly given by Fg(!!0) =Mz totrBz: (10) Using G 1(!)determined above, we can calculate how the effective force Sg(!)rBz(!)(see Eq. 2) depends on fre- quency. Fig. 6 shows that for !!0one obtains the expectedresultSg(!!0) =ijMz tot, while a strong frequency de- pendence sets in above the magnon gap, for !&!p. This is the precession frequency of spins in the external magnetic field. In general, both the screening of forces (parametrized bySg(!)) and the internal dynamics (described by G 1(!)) determines the response of skyrmions, V(!) = G(!)Sg(!)rBz(!). Therefore it is in general not possi- ble to extract, e.g., the mass of the skyrmion as described by G 1(!)from just a measurement of the response to field gra- dients. It is, however, instructive to ask what “apparent” mass one obtains, when the frequency dependence of Sg(!)is ig- nored. We therefore define the “apparent” dynamics G 1 a(!) byGa(!)Sg(!= 0) = G(!)Sg(!). The matrix elements ofG 1 a(!)are shown in Fig. 3 as dashed lines. The appar- ent mass for gradient-driven motion, for example, turns out to be more than a factor of three smaller then the value ob- tained from the diffusive motion clearly showing the impor- tance of screening effects on external forces. The situation is even more dramatic when skyrmions are driven by electric currents. V . CURRENT-DRIVEN MOTION Currents affect the motion of spins both via adiabatic and non-adiabatic spin torques30. Therefore one obtains two types of forces on the spin texture even in the static limit19–22,24. The effect of a time-dependent, spin-polarized current on the magnetic texture can be modelled by supplementing the right hand side of eq. (3) with a spin torque term TST, TST= (vsr)Mr+ M[Mr(vsr)Mr]:(11) The first term is called the spin-transfer-torque term and is derived under the assumption of adiabaticity: the conduction- electrons adjust their spin orientation as they traverse the mag- netic sample such that it points parallel to the local magnetic moment Mrowing toJHandJsd. This assumptions is justi- fied as the skyrmions are rather large smooth objects (due to the weakness of spin-orbit coupling). The second so called - term describes the dissipative coupling between conduction- electrons and magnetic moments due to non-adiabatic effects. Bothandare small dimensionless constants in typical ma- terials. From the Thiele approach one obtains the force Fc(!!0) =Gvs+Dvs: (12) For a Galilei-invariant system one obtains =. In this special limit, one can easily show that an exact solution of the LLG equations in the presence of a time-dependent current, described by vs(t)is given by M(r Rt 1vs(t0)dt0)pro- vided, M(~ r)is a static solution of the LLG equation for vs= 0. This implies that for =, the skyrmion motion exactly follows the external current, _R(t) =vs(t). Using Eq. (2), this implies that for =one has G 1(!) =Sc(!). Defin- ing the apparent dynamics, as above, Ga(!)Sc(!= 0) = G(!)Sc(!)one obtains a frequency independent G 1 a(!) =6 0 1 2 3 4 5 ω / ωp-4 π-10-50β D(0)510Re/Im ScIm Sc11 Im Sc21Re Sc11 Re Sc21 FIG. 7: Dynamical coupling coefficients (symbols) for the current- driven motion ( = 0:2,= 0:1,J= 1,= 0:18J,B= 0:0278 ). These curves follow almost the corresponding matrix elements of G 1(!)shown as dashed lines. A deviation of symbols and dashed line is only sizable for Re S11 c. 0 1 2 3 4 5 ω/ωp-5051015 mα=0.2,β=0 α=0.2,β=0.1 0 1 2 3 4 5 ω/ωp0510 α Γα=0.2,β=0.15 α=0.2,β=0.19 α=0.2,β=0.3 FIG. 8: Mass m(!)and gyrodamping (!)as functions of the driving frequency !for the current-driven motion. Note that both M and vanish for=. Sc(!= 0) =D 1 iyG: the apparent effective mass and gyrodamping are exactly zero in this limit and the skyrmion follows the current without any retardation. For 6=, the LLG equations predict a finite apparent mass. Numerically, we find only very small apparent masses, ma c/ , see dot-dashed line in upper-right panel of Fig. 3, where the case = 0:2,= 0:1is shown. This is anticipated from the anal- ysis of the=case: As the mass vanishes for == 0, it will be small as long as both andare small. Indeed even for6=this relation holds approximately as shown in Fig. 7. The only sizable deviation is observed for Re S11 c for which the Thiele equation predicts Re S11 c(!!0) =D while Re G 111(!!0) =Das observed numerically. A better way to quantify that the skyrmion follows the cur-rent even for 6=almost instantaneously is to calculate the apparent mass and gyrodamping for current driven mo- tion, where only results for = 0:2and= 0:1have been shown. As these quantities vanish for =, one can ex- pect that they are proportional to at least for small ;. This is indeed approximately valid at least for small frequen- cies as can be seen from Fig. 8. Interestingly, one can even obtain negative values for > (without violating causal- ity). Most importantly, despite the rather large values for andused in our analysis, the apparent effective mass and gyrodamping remain small compared to the large values ob- tained for force-driven motion or the intrinsic dynamics. This shows that retardation effects remain tiny when skyrmions are controlled by currents. VI. CONCLUSIONS In conclusion, we have shown that skyrmions in chiral mag- nets are characterised by a number of unique dynamical prop- erties which are not easily found in other systems. First, their damping is small despite the fact that skyrmions are large composite objects. Second, despite the small damping, the diffusion constant remains small. Third, despite a huge iner- tial mass, skyrmions react almost instantaneously to external currents. The combination of these three features can become the basis for a very precise control of skyrmions by time- dependent currents. Our analysis of the skyrmion motion is based on a two- dimensional model where only a single magnetic layer was considered. All qualitative results can, however, easily be generalized to a film with NLlayers. In this case, all terms in Eq. (1) get approximately multiplied by a factor NLwith the exception of the last term, the random force, which is en- hanced only by a factorpNL. As a consequence, the diffu- sive motion is further suppressed by a factor 1=pNLwhile the current- and force-driven motion are approximately unaf- fected. An unexpected feature of the skyrmion motion is the an- tidamping arising from the gyrodamping. The presence of antidamping is closely related to another important property of the system: both the dynamics of the skyrmion and the ef- fective forces acting on the skyrmion are strongly frequency dependent. In general, in any device based on skyrmions a combination of effects will play a role. Thermal fluctuations are always present in room-temperature devices, the shape of the device will exert forces13,14and, finally, we have identified the cur- rent as the ideal driving mechanism. In the linear regime, the corresponding forces are additive. The study of non-linear effects and the interaction of several skyrmions will be impor- tant for the design of logical elements based on skyrmions and this is left for future works. As in our study, we expect that dynamical screening will be important in this regime.7 30 40 50 60 70 30 40 50 60 70 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0.004 FIG. 9: Skyrmion density based on the normalized z-component of the magnetization. Acknowledgments The authors are greatful for insightful discussions with K. Everschor and Markus Garst. Part of this work was funded through the Institutional Strategy of the University of Cologne within the German Excellence Initiative” and the BCGS. C.S. thanks the University of Tokyo for hospitality during his re- search internship where part of this work has been performed. N.N. was supported by Grant-in-Aids for Scientific Research (No. 24224009) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan, and by the Strategic International Cooperative Program (Joint Research Type) from Japan Science and Technology Agency. J.I. is sup- ported by Grant-in-Aids for JSPS Fellows (No. 2610547). Appendix A: Definition of the Skyrmion’s centre coordinate In order to calculate the Green’s function, Eq. (3), one needs to calculate the velocity-velocity correlation function. Therefore it is necessary to track the skyrmion position throughout the simulation. Mostly two methods have been used so far for this25: (i) tracking the centre of the topological charge and (ii) tracking the core of the Skyrmion (reversal of magnetization). The topological charge density top(r) =1 4^ n(r)(@x^ n(r)@y^ n(r)) (A1) integrates to the number of Skyrmions in the system. There- fore for our case of a single Skyrmion in the ferromagnetic background this quantity is normalized to 1. The center of topological charge can therefore be defined as R=Z d2rtop(r)r (A2) For the case of finite temperature this method can, however, not be used directly. Thermal fluctuations in the ferromagnetic background far away from the skyrmion lead to a large noise to this quantity which diverges in the thermodynamic limit. A similar problem arises when tracking the center using the magnetization of the skyrmion.One therefore needs a method which focuses only on the region close to the skyrmion center. To locate the skyrmion, we use thez-component of the magnetization but take into ac- count only points where Mz(r)< |